Synthesis and self-assembly of multiple thermoresponsive ...

173
Synthesis and Self-Assembly of Multiple Thermoresponsive Amphiphilic Block Copolymers A dissertation submitted to Potsdam 2011 Presented by University of Potsdam for the degree of “doctor rerum naturalium” (Dr. rer. nat.) in Macromolecular Chemistry Jan Weiß Department of Chemistry Supervisor Prof. A. Laschewsky

Transcript of Synthesis and self-assembly of multiple thermoresponsive ...

  • Synthesis and Self-Assembly of MultipleThermoresponsive Amphiphilic Block Copolymers

    A dissertation submitted to

    Potsdam 2011

    Presented by

    University of Potsdam

    for the degree of

    doctor rerum naturalium

    (Dr. rer. nat.)

    in Macromolecular Chemistry

    Jan Wei

    Department of Chemistry

    Supervisor Prof. A. Laschewsky

  • Published online at the Institutional Repository of the University of Potsdam: URL http://opus.kobv.de/ubp/volltexte/2011/5336/ URN urn:nbn:de:kobv:517-opus-53360 http://nbn-resolving.de/urn:nbn:de:kobv:517-opus-53360

  • Danksagung

    Mein Dank geht an Prof. Dr. A. Laschewsky fr die Mglichkeit meine Dok-

    torarbeit in seiner Gruppe anfertigen zu knnen und fr die Untersttzung

    und Diskussionen whrend dieser Zeit. Ferner mchte ich Dr. J. van Heijst

    und Dr. H. Regger, der das Ende dieser Arbeit aufgrund tragischer Um-

    stnde leider nicht mehr miterleben kann, fr die Mglichkeit danken, tem-

    peraturabhngige NMR Spektren an der ETH Zrich messen zu knnen,

    Dr. C. Bttcher fr die TEM Messungen, Prof. Dr. S. Beuermann und Herrn

    S. Prentzel fr SEC Messungen, Dr. N. Hildebrandt und Dr. S. Stufler fr

    den Zugang zum Fluoreszenzspektrometer sowie Prof. Dr. J. Beckmann und

    Prof. Dr. H. Frauenrath fr hilfreiche Diskussionen.

    Ausserdem mchte ich Dr. Achille Bivigou-Koumba fr die zahlreichen

    Diskussionen und die praktischen Ratschlge, besonders in der An-

    fangsphase meiner Promotionszeit und Maike Lukowiak fr den Zugang zu

    einiger Literatur danken.

    Des Weiteren mchte ich Dr. Nezha Badi fr die hilfreichen Diskussionen,

    Korrekturen dieser Arbeit, die Untersttzung sowie fr die wundervolle Zeit

    auch neben der Arbeit, die meine drei Jahre in Potsdam unvergesslich macht,

    von ganzem Herzen danken.

    Zu guter Letzt geht noch ein besonderer Dank an meine Familie, insbeson-

    dere an meine Mutter fr ihre fortwhrende Untersttzung in jeglicher Hin-

    sicht.

  • List of Publications

    Parts of this Ph. D. thesis were published in scientific journals and were presented in oral

    presentations as well as on posters at several occasions by the author of this thesis.

    Publications (peer-reviewed)

    [1] Universal Polymer Analysis by 1H NMR Using Complementary Trimethylsilyl End

    Groups, M. Pch, D. Zehm, M. Lange, I. Dambowsky, J. Weiss, A. Laschewsky, J. Am.

    Chem. Soc., 2010, 132, 8757-8765.

    [2] Self-Assembly of Double Thermoresponsive Block Copolymers End-capped with

    Complementary Trimethylsilyl Groups, J. Weiss, C. Bttcher, A. Laschewsky, Soft Mat-

    ter, 2011, 7, 483-492.

    [3] Temperature Induced Self-Assembly of Triple Thermoresponsive Triblock Copoly-

    mers in Aqueous Solution, J. Weiss, A. Laschewsky, Langmuir, 2011, 27, 4465-4473.

    [4] Ferrocenyl (Meth)Acrylates in RAFT Polymerization, C. Herfurth, D. Voll, J. Buller,

    J. Weiss, C. Barner-Kowollik, A. Laschewsky, submitted to J. Polym. Sci. Part A: Polym.

    Chem.

    Publications (to be submitted)

    [5] Water-soluble Random and Alternating Copolymers of Styrene Monomers with

    Adjustable Lower Critical Solution Temperature, J. Weiss, A. Li, E. Wischerhoff,

    A. Laschewsky*, to be submitted.

    [6] Facile One-Step Synthesis of Double Thermosensitive Diblock Copolymers, J. Weiss,

    A. Laschewsky*, to be submitted.

    IV

  • Oral Presentations

    [1] A Trimethylsilyl-labeled RAFT-Agent as NMR Probe for Reversible Block Copolymer

    Self-Assembly, 13.07.2010, 43rd IUPAC World Polymer Congress, Glasgow, UK.

    [2] Multistep Self-Assembly of Thermoresponsive Block Copolymer Surfactants,

    03.05.2011, 45th International Detergency Conference (IDC), Dsseldorf, Germany.

    Poster Contributions

    [1] Temperature Induced Self-Assembly of Triple-Responsive Triblock Copolymers in

    Aqueous Solutions, PhD Students Workshop Functional Soft Matter, Potsdam, Ger-

    many, 2010.

    [2] Temperature Induced Self-Assembly of Triple-Responsive Triblock Copolymers in Di-

    lute Aqueous Solution, 43rd IUPAC World Polymer Congress, Glasgow, UK, 2010.

    [3] Self-Assembly of Triple-Thermoresponsive Triblock Copolymers in Dilute Aqueous

    Solution, Polymers in Biomedicine and Electronics - Biannual Meeting of the GDCh-

    Division Macromolecular Chemistry and Polydays 2010, Berlin, Germany, 2010.

    [4] Improving the IQ of Intelligent Block Copolymer Surfactants: Designs for Multiple

    Switching, 6th European Detergency Conference, Fulda, Germany, 2010.

    [5] Sequential Self-Assembly of Multiple Thermoresponsive Block Copolymers, Makro-

    molekulares Kolloquium, Freiburg, Germany, 2011.

    V

  • Summary

    In the present thesis, the self-assembly of multi thermoresponsive block copolymers in di-

    lute aqueous solution was investigated by a combination of turbidimetry, dynamic light

    scattering, TEM measurements, NMR as well as fluorescence spectroscopy. The succes-

    sive conversion of such block copolymers from a hydrophilic into a hydrophobic state in-

    cludes intermediate amphiphilic states with a variable hydrophilic-to-lipophilic balance.

    As a result, the self-organization is not following an all-or-none principle but a multistep

    aggregation in dilute solution was observed. The synthesis of double thermoresponsive di-

    block copolymers as well as triple thermoresponsive triblock copolymers was realized us-

    ing twofold-TMS labeled RAFT agents which provide direct information about the average

    molar mass as well as residual end group functionality from a routine 1H NMR spectrum.

    First a set of double thermosensitive diblock copolymers poly(N-n-propylacrylamide)-b-

    poly(N-ethylacrylamide) was synthesized which differed only in the relative size of the two

    blocks. Depending on the relative block lengths, different aggregation pathways were found.

    Furthermore, the complementary TMS-labeled end groups served as NMR-probes for the

    self-assembly of these diblock copolymers in dilute solution. Reversible, temperature sensi-

    tive peak splitting of the TMS-signals in NMR spectroscopy was indicative for the formation

    of mixed star-/flower-like micelles in some cases.

    T1 T2n m

    Moreover, triple thermoresponsive triblock copolymers from poly(N-n-

    propylacrylamide) (A), poly(methoxydiethylene glycol acrylate) (B) and poly(N-

    ethylacrylamide) (C) were obtained from sequential RAFT polymerization in all possible

    block sequences (ABC, BAC, ACB). Their self-organization behavior in dilute aqueous

    solution was found to be rather complex and dependent on the positioning of the different

    blocks within the terpolymers. Especially the localization of the low-LCST block (A) had a

    large influence on the aggregation behavior. Above the first cloud point, aggregates were

    VI

  • only observed when the A block was located at one terminus. Once placed in the middle,

    unimolecular micelles were observed which showed aggregation only above the second

    phase transition temperature of the B block. Carrier abilities of such triple thermosensitive

    triblock copolymers tested in fluorescence spectroscopy, using the solvatochromic dye Nile

    Red, suggested that the hydrophobic probe is less efficiently incorporated by the polymer

    with the BAC sequence as compared to ABC or ACB polymers above the first phase transition

    temperature.

    In addition, due to the problem of increasing loss of end group functionality during the

    subsequent polymerization steps, a novel concept for the one-step synthesis of multi

    thermoresponsive block copolymers was developed. This allowed to synthesize double

    thermoresponsive di- and triblock copolymers in a single polymerization step. The copoly-

    merization of different N-substituted maleimides with a thermosensitive styrene derivative

    (4-vinylbenzyl methoxytetrakis(oxyethylene) ether) led to alternating copolymers with vari-

    able LCST. Consequently, an excess of this styrene-based monomer allowed the synthesis of

    double thermoresponsive tapered block copolymers in a single polymerization step.

    N OO

    R

    t0

    RAFT or ATRP

    OO

    4

    alternating block homopolymer

    Furthermore, by using bifunctional initiators, even double thermosensitive binary tri-

    block copolymers could be synthesized. Both types of polymers showed an aggregation be-

    havior similar to the one of block copolymers obtained by the classical step-wise approach

    indicating the successful one-step synthesis of multi responsive block copolymers.

    VII

  • Zusammenfassung

    Im Rahmen der vorliegenden Arbeit wurde die Selbstorganisation von mehrfach ther-

    misch schaltbaren Blockcopolymeren in verdnnter wssriger Lsung mittels Trbungspho-

    tometer, dynamischer Lichtstreuung, TEM Messungen, NMR sowie Fluoreszenzspek-

    troskopie untersucht. Die schrittweise berfhrung eines hydrophilen in ein hydrophobes

    Blockcopolymer beinhaltet ein oder mehr amphiphile Zwischenstufen mit einstellbarem

    hydrophilen zu lipophilen Anteil (HLB). Dies fhrt dazu, dass die Selbstorganisation

    solcher Polymere in Lsung nicht nur einem Alles-oder-nichts-Prinzip folgt sondern ein

    mehrstufiges Aggregationsverhalten beobachtet wird. Die Synthese von doppelt thermisch

    schaltbaren Diblockcopolymeren und dreifach thermisch schaltbaren Triblockcopolymeren

    wurde durch sequenzielle RAFT Polymerisation realisiert. Dazu wurden zweifach TMS-

    markierte RAFT Agentien verwendet, welche die Bestimmung der molaren Masse sowie der

    verbliebenen Endgruppenfunktionalitt direkt aus einem 1H NMR Spektrum erlauben. Mit

    diesen RAFT Agentien wurde zunchst eine Serie von doppelt thermisch schaltbaren Di-

    blockcopolymeren aus Poly(N-n-propylacrylamid)-b-Poly(N-ethylacrylamid), welche sich

    lediglich durch die relativen Blocklngen unterscheiden, hergestellt. In Abhngigkeit von der

    relativen Blocklnge wurde ein unterschiedliches Aggregationsverhalten der Diblockcopoly-

    mere in verdnnter wssriger Lsung beobachtet. Des Weiteren wirken die komplemen-

    tr TMS-markierten Endgruppen als NMR-Sonden whrend der schrittweisen Aggregation

    dieser Polymere. Reversible, temperaturabhngige Peakaufspaltung der TMS-Signale in der

    NMR Spektroskopie spricht fr eine Aggregation in gemischte stern-/blumenartige Mizellen,

    in denen ein Teil der hydrophoben Endgruppen in den hyrophoben Kern zurckfaltet.

    T1 T2n m

    Obendrein wurden dreifach thermisch schalbare Triblockcopolymere aus Poly(N-n-

    propylacrylamid) (A), Poly(methoxydiethylen glycol acrylat) (B) und Poly(N-ethylacrylamid)

    (C) in allen mglichen Blocksequenzen (ABC, BAC, ACB) durch schrittweisen Aufbau mit-

    tels RAFT Polymerisation erhalten. Das Aggregationsverhalten dieser Polymere in verdnn-

    VIII

  • ter wssriger Lsung war relativ komplex und hing stark von der Position der einzelnen

    Blcke in den Triblockcopolymeren ab. Besonders die Position des Blocks mit der niedrig-

    sten LCST (A) war ausschlaggebend fr die resultierenden Aggregate. So wurde oberhalb der

    ersten Phasenbergangstemperatur nur Aggregation der Triblockcopolymere beobachtet,

    wenn der A Block an einem der beiden Enden der Polymere lokalisiert war. Wurde der A

    Block hingegen in der Mitte der Polymere positioniert, entstanden unimere Mizellen zwis-

    chen der ersten und zweiten Phasenbergangstemperatur, welche erst aggregierten, nach-

    dem der zweite Block (B) seinen Phasenbergang durchlief. Die Transportereigenschaften

    dieser Triblockcopolymere wurden mittels Fluoreszenzspektroskopie getestet. Dazu wurde

    die Einlagerung eines hydrophoben, solvatochromen Fluoreszenzfarbstoffes, Nilrot, in Ab-

    hngigkeit der Temperatur untersucht. Im Gegensatz zu den Polymeren mit der Blockse-

    quenz ABC oder ACB, zeigten die Polymere mit der Sequenz BAC eine verminderte Aufnah-

    mefhigkeit des hydrophoben Farbstoffes oberhalb des ersten Phasenbergangs, was auf die

    fehlende Aggregation und die damit verbundenen relativ kleinen hydrophoben Domnen

    der unimolekularen Mizellen zwischen der ersten und zweiten Phasenbergangstemperatur

    zurckzufhren ist.

    Aufgrund des zunehmenden Verlustes von funktionellen Endgruppen whrend der RAFT

    Synthese von Triblockcopolymeren wurde ein neuartiges Konzept zur Einschrittsynthese

    von mehrfach schaltbaren Blockcopolymeren entwickelt. Dieses erlaubt die Synthese von

    mehrfach schaltbaren Diblock- und Triblockcopoylmeren in einem einzelnen Reaktions-

    schritt. Die Copolymeriation von verschiedenen N-substituierten Maleimiden mit einem

    thermisch schaltbaren Styrolderivat (4-Vinylbenzylmethoxytetrakis(oxyethylene) ether) er-

    gab alternierende Copolymere mit variabler LCST. Die Verwendung eines berschusses

    dieses styrolbasierten Monomers erlaubt ferner die Synthese von Gradientenblockcopoly-

    meren in einem einzelnen Polymerisationsschritt.

    N OO

    R

    t0

    RAFT or ATRP

    OO

    4

    alternating block homopolymer

    IX

  • Bifunktionelle Initiatoren ergaben, dem gleichen Reaktionsprinzip folgend, doppelt

    schaltbare binre Triblockcopolymere. Die so hergestellten Blockcopolymere zeigten ein

    vergleichbares Aggregationsverhalten in verdnnter wssriger Lsung wie Blockcopolymere,

    die durch klassiche sequenzielle kontrolliert radikalische Polymerisation erhalten werden.

    X

  • List of Abbreviations

    AGET activator generated by electron transfer

    anal. analysis

    ARGET activators regenerated by electron transfer

    ATRP atom transfer radical polymerization

    b broad signal (NMR)

    calcd calculated

    CP cloud point

    CRP controlled radical polymerization

    CTA chain transfer agent

    d doublet (NMR)

    chemical shift (NMR)

    DCM dichloromethane

    DBPO dibenzoyl peroxide

    DIEA diisopropylethylamine

    DIPA diisopropylamine

    DMF N,N-dimethylformamide

    DMSO dimethylsulfoxide

    DP average degree of polymerization

    EI electron impact ionization

    eq. equivalents

    ESI electrospray ionization

    FRP free radical polymerization

    GPC gel permeation chromatography1H NMR proton nuclear magnetic resonance13C NMR 13carbon nuclear magnetic resonance

    HMDS hexamethyldisilazane

    HRMS high resolution mass spectrometry

    HV high vacuum

    IR infra red

    J J coupling constant (NMR)

    LFRP living free radical polymerization

    m multiplet (NMR)

    XI

  • M molar

    [M]+ molecular ion (MS)

    MDEGA methoxydiethylene glycol acrylate

    Mn number average molecular weight

    NDM N-decylmaleimide

    NEA N-ethylacrylamide

    NMM N-methylmaleimide

    NMP nitroxide mediated polymerization

    NMR nuclear magnetic resonance

    NPA N-propylacrylamide

    NPM N-propylmaleimide

    NPEGM N-PEG750-maleimide

    NSM N-ethylthiomethyl maleimide

    NSOM N-ethylsulfoxymethyl maleimide

    NtBAlaM N,N-maleoyl-L-alanine tert.-butylester

    NtBGlyM N,N-maleoyl-L-glycine tert.-butylester

    NTESM N-(3-triethylsilyl)propargyl maleimide

    NTMSM N-(3-trimethylsilyl)propyl maleimide

    PDI polydispersity index

    PEG poly(ethylene glycol)

    PEO poly(ethylene oxide)

    ppm parts per million (NMR)

    PS polystyrene

    RAFT reversible addition fragmentation chain transfer

    Rf retention or retardation factor (thin layer chromatography)

    s singlet (NMR)

    t triplet (NMR)

    TBAF tetrabutylammonium fluoride

    TEA triethylamine

    TEM transmission electron microscopy

    TEMPO 2,2,6,6-tetramethylpiperidinyloxyl

    TES triethylsilyl

    TFA trifluoroacetic acid

    THF tetrahydrofuran

    TLC thin layer chromatography

    XII

  • TMS trimethylsilyl

    TMSR TMS group on the R group of a TMS-labeled RAFT agent

    TMSZ TMS group on the Z group of a TMS-labeled RAFT agent

    TEGDME triethyleneglycol dimethylether

    UV ultraviolet

    VBTOE 4-vinylbenzyl methoxytetrakis(oxyethylene) ether

    XIII

  • Contents

    1 Scope and Motivation 1

    2 Introduction 3

    2.1 Controlled Radical Polymerization Techniques . . . . . . . . . . . . . . . . . . . 3

    2.1.1 Nitroxide Mediated Radical Polymerization . . . . . . . . . . . . . . . . . 4

    2.1.2 Atom Transfer Radical Polymerization . . . . . . . . . . . . . . . . . . . . . 6

    2.1.3 Reversible Addition-Fragmentation Chain Transfer Polymerization . . . 8

    2.2 Block Copolymer Self-Assembly . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

    2.2.1 Stimuli Responsive Block Copolymers . . . . . . . . . . . . . . . . . . . . . 19

    2.3 Alternating Copolymerization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

    2.3.1 Alternating Copolymers of Maleic Anhydride and Styrene . . . . . . . . . 31

    2.3.2 Copolymerization of N-substituted Maleimides with Styrene . . . . . . . 33

    2.3.3 One-Step Synthesis of Diblock Copolymers . . . . . . . . . . . . . . . . . 34

    2.4 Previous Investigations at Potsdam University . . . . . . . . . . . . . . . . . . . . 35

    2.4.1 Twofold TMS-Labeled RAFT-Agents . . . . . . . . . . . . . . . . . . . . . . 35

    2.4.2 Mono and Double Responsive Triblock Copolymers . . . . . . . . . . . . 38

    2.5 Objectives of this Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

    3 Results and Discussion 43

    3.1 Synthesis of RAFT-Agents and Monomers . . . . . . . . . . . . . . . . . . . . . . . 43

    3.1.1 Synthesis of RAFT-Agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

    3.1.2 Synthesis of Monomers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

    3.2 TMS-Labeled RAFT-Agents and Styrenes . . . . . . . . . . . . . . . . . . . . . . . 45

    3.3 Sequential RAFT-Synthesis of Multiple Thermoresponsive Block Copolymers . 47

    3.3.1 Synthesis and Solution Properties of Homopolymers . . . . . . . . . . . . 48

    3.3.2 Double Thermoresponsive Diblock Copolymers . . . . . . . . . . . . . . . 49

    3.3.3 Triple Thermoresponsive Triblock Copolymers . . . . . . . . . . . . . . . 68

    XV

  • 3.4 One-Step Synthesis of Multi Responsive Block Copolymers . . . . . . . . . . . . 85

    3.4.1 Synthesis of Monomers and ATRP Initiators . . . . . . . . . . . . . . . . . 85

    3.4.2 Homopolymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

    3.4.3 Alternating Copolymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

    3.4.4 Diblock Copolymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

    3.4.5 ABA and BAB Triblock Copolymers . . . . . . . . . . . . . . . . . . . . . . 105

    4 Conclusions 111

    5 Experimental 115

    5.1 Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

    5.2 General Procedures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

    5.3 RAFT-Agents and ATRP Initiators . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

    5.4 Monomers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

    5.4.1 Acrylamides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

    5.4.2 Acrylates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

    5.4.3 Styrenics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

    5.4.4 Maleimides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

    6 References 129

    7 Appendix 143

    XVI

  • 1 Scope and Motivation

    Hierarchically structured nanomaterials, as often present in biological systems, are typi-

    cally obtained using a combined bottom-up and top-down approach.1 The top-down

    approach, although still dominating in todays industries, will face increasing technologi-

    cal challenges and might reach its physical limitations in the near future. By contrast, the

    bottom-up approach aims to build up ordered structures from (macro)molecular precur-

    sors, attempting to benefit from their built-in information for higher structure formation.

    The propensity to hierarchical structure formation, thus, is programmed on the molecu-

    lar level and translated into building blocks with a defined shape on the supramolecular

    level. Accordingly, the self-organization of appropriate macromolecular building blocks is a

    promising bottom-up route toward well-defined nanomaterials.24

    Within this context, amphiphilic block copolymers in aqueous media represent an inter-

    esting class of self-organizing molecules since they are known to form a variety of aggre-

    gates such as spherical micelles, worm-like (i.e. cylindrical) micelles, as well as vesicles

    in dilute aqueous solution.510 Especially, stimuli responsive polymers, also often referred

    to as smart, found increasing attention in the last decade due to their potential applica-

    tion in industry ranging from biomedical to material science.6 The ability to control the

    aggregation behavior of stimuli-sensitive polymers may help to address future challenges in

    nanoscience and provides fundamental knowledge for the design of smart materials with

    tunable properties. This thesis aims at understanding how multi responsive block copoly-

    mers self-assemble in aqueous solution and tries to develop solutions for current limitations

    in polymer synthesis and characterization as well as investigation of the formed aggregates

    on a molecular level.

    1

  • 2 Introduction

    2.1 Controlled Radical Polymerization Techniques

    Free radical polymerization (FRP) is widely used in order to obtain high molecular weight

    polymers due to (i) the compatibility with a variety of monomers, such as (meth)acrylates,

    (meth)acrylamides, styrenics, dienes as well as other vinylic monomers; (ii) its tolerance

    against many functionalities within monomers and solvent, e.g., OH, NR2, COOH, CONR2,

    SO3; (iii) its high compatibility with various reaction conditions (e.g., bulk, solution, emul-

    sion, mini-emulsion, and suspension); (iv) its relatively low costs compared to other tech-

    nologies. Although widely used in industry and research laboratories, FRP has a significant

    drawback. The variety of possible termination reactions result in limited control over mo-

    lecular weight, molecular weight distribution as well as end groups and architecture of the

    desired polymers.11 Moreover, block copolymers with defined block sequences and prede-

    termined relative block lengths are practically not accessible by conventional free radical

    polymerization. In order to overcome these limitations and enable the synthesis of well de-

    fined block copolymers, several controlled radical polymerization (CRP) techniques were

    invented within the past two decades, such as nitroxide mediated polymerization12 (NMP),

    atom transfer radical polymerization13 (ATRP), or reversible addition-fragmentation chain

    transfer polymerization14 (RAFT). Before the discovery of the CRP principle, living ionic

    polymerizations were one of the few available tools to achieve control over molecular weight

    and architecture with low polydispersities. The major disadvantages of ionic polymeriza-

    tions, however, are the very stringent reaction conditions, typically the complete absence of

    oxygen and water, as well as the sensitivity to most functional groups.15 The development

    of controlled radical polymerization techniques combined the best of both attempts, com-

    patibility with a wide range of functional monomers, low polydispersities, and control over

    molecular weight and architecture. The key principle of controlled radical polymerizations

    is based on reversible chain termination (Scheme 1).

    3

  • Nitroxide Mediated Radical Polymerization

    Polymer R Polymer R

    Scheme 1: Basic principle of Controlled Radical Polymerizations: Reversible interchange between

    dormant and active chain ends.

    As a result, the polymer chains can grow simultaneously throughout the reaction with

    very low concentrations of free radicals present at all times during the polymerization pro-

    cess, thus minimizing irreversible termination reactions, such as recombination and dispro-

    portionation. However, controlled radical polymerization techniques are no true living

    systems1619 and termination reactions are not negligible. Nevertheless, they display impor-

    tant characteristics of a living polymerization such as (i) a linear relation between molecu-

    lar weight and conversion, (ii) molecular weight equal or close to the theoretical molecular

    weight, (iii) low polydispersity, and (iv) defined end groups. For this reason, they are some-

    times referred to as living free radical polymerization (LFRP) in the literature.

    2.1.1 Nitroxide Mediated Radical Polymerization

    Nitroxide mediated radical polymerization resulted from intentions of the australian Com-

    monwealth Scientific and Industrial Research Organization (CSIRO) team around D. H.

    Solomon to study the initiation of free radical polymerizations by trapping the propaga-

    ting radical.20 The trapping agent used, 2,2,6,6-tetramethylpiperidinyloxy (TEMPO), does

    not react with heteroatom-centered radicals but adds to carbon-centered radicals with near

    diffusion-controlled rates.21 The fact that trapped radicals with more than one monomer

    unit were found, in combination with the observed thermal instability of the formed

    alkoxyamines, led to the assumption that thermal dissociation and reversible trapping can

    lead to the controlled formation of oligomeric compounds.

    Based on this initial work of Solomon, Rizzardo and Moad, low molecular weight polymers

    and oligomers were obtained at temperatures of 80-100C.22 Low polydispersity polystyrene

    was synthesized by further increasing the temperature to 130C in bulk using dibenzoyl per-

    oxide (DBPO) as initiator and TEMPO as mediator.23 Since then ongoing research led to

    4

  • Nitroxide Mediated Radical Polymerization

    improved nitroxides12 which allow polymerization of many different monomer classes, e. g.,

    acrylates, acrylamides, 1,3-dienes as well as acrylonitrile based monomers.

    The polymerization mechanism of the nitroxide mediated radical polymerization is based on

    a kinetic phenomenon called persistent radical effect24 (Scheme 2). After heterolytic cleavage

    of the initiator into the initiating radical X and the mediating radical R, small amounts of

    the initiating radical will undergo radical-radical coupling. The mediating radical R, how-

    ever, cannot undergo homocoupling which results in an overall excess of mediating radicals

    compared to initiating/propagating radicals.

    X R X

    X X

    Coupling

    + R

    M

    X (M)n R

    Polymerization

    Scheme 2: Principle scheme for the nitroxide mediated radical polymerization.

    The increasing efficiency of the formation of dormant chain ends with increasing excess

    of mediating radicals regulates this equilibrium. The resulting small quantity of propagating

    radicals with a large overall excess of dormant polymer chains gives rise to the persistent ra-

    dical effect (PRE) and, hence, potential control over molecular weight and molecular weight

    distribution. The upper limit for controlled molecular weight lays at 150000-200000 g mol1

    for NMP polymerizations.25

    5

  • Atom Transfer Radical Polymerization

    2.1.2 Atom Transfer Radical Polymerization

    The best established controlled radical polymerization technique to date is the transition

    metal catalyzed atom transfer radical polymerization (ATRP) which was independently dis-

    covered by Sawamoto et al. 26 and Matyjaszewski et al. 27 in 1995. The mechanism of ATRP

    (Scheme 3) is based on a reversible redox process which generates the active species, usually

    a CuI-complex. The oxidized species leads to a reversible abstraction of a halogen atom (Cl,

    Br) from the dormant species R-X to generate the active radical R.

    X R + Mtn Y/Ligandkact

    kdeact

    R

    kp

    Monomer

    Termination

    + Mtn-1 Y/LigandX

    Scheme 3: General mechanism of the atom transfer radical polymerization.

    Molecular weights up to 150000 g mol1 have been successfully achieved by ATRP. How-

    ever, at higher molecular weights increasing amounts of termination reactions suggest an

    upper limit for controlled radical polymer synthesis.28 Such termination reactions mainly

    include recombination and disproportionation and result from interactions of CuII species

    with both the growing radical as well as its dormant species. In contrast to the normal

    ATRP, where the initiating radicals are formed from an alkyl-halide and a transition metal

    in its lower oxidation state, reverse ATRP uses classical initiators (e. g. AIBN) in combination

    with a transition metal-complex in its higher oxidation state.2931 Various monomer types

    have been polymerized by ATRP such as styrenes, (meth-)acrylates, (meth-)acrylamides and

    others.13 However, some monomer classes are not accessible such as monomers contain-

    ing acidic side chains, since they can protonate the ligands which form the corresponding

    carboxylates, as well as halogenated alkenes, alkyl-substituted olefines, and vinyl esters. In

    ATRP, likewise to other controlled polymerization techniques, the initiator should ensure a

    fast initiation step compared to propagation.

    6

  • Atom Transfer Radical Polymerization

    Therefore, some general considerations for the initiator design should be taken into

    account:

    The initiator quality decreases from tertiary to secondary to primary alkyl halides and

    addition of stabilizing groups, with relative efficiencies in the order CN > C(O)R >

    C(O)OR > Ph > Cl > Me, will improve the initiator quality.3234

    Although the bond strength for alkyl halides decreases in the order R-Cl > R-Br > R-I

    and, hence, alkyl iodides should be the most efficient ATRP initiators, their use requires

    certain precautions. For instance, their light sensitivity can lead to the formation of

    metal iodide complexes (e. g., CuI2 which is thermodynamically unstable) or hetero-

    lytic cleavage resulting in degenerative transfer reactions.35 Accordingly, bromides

    and chlorides are preferentially used. Also pseudohalogens such as thiocyanates (SCN)

    have been used36 but showed slow initiation for both styrene and methacrylate (MA).

    The choice of catalyst may also influence the initiation efficiency. For example,

    2-bromoisobutyrophenone initiates the controlled polymerization of methylmeth-

    acrylate (MMA) when ruthenium or nickel complexes are used but turns out to be

    useless in the copper-mediated ATRP. This is ascribed to the reduction of the resulting

    electrophilic radical by the CuI species due to the lower redox potentials of copper-

    based catalysts.

    The order of reagent addition might be important. Upon slow addition of the catalyst

    to the initiator/monomer solution the rate of termination during the initiation period

    was found to be reduced.13

    Beside the initiator and transition metal catalyst, the ligand is very important. Its major

    role is to provide solubility for the transition-metal salt in the reaction medium and to adjust

    the redox potential of the metal center. For the copper-mediated ATRP, nitrogen-containing

    chelating ligands have predominantly been used not least because of the much lower effi-

    ciency of sulfur, oxygen or phosphorus ligands due to unfavorable binding constants and/or

    improper electronic effects. In principle, the activity of an ATRP ligand decreases with de-

    creasing nitrogen atoms and increasing linking carbon atoms.37 The major disadvantage of

    ATRP is, beside the relatively high temperatures usually used, the high amounts of (gener-

    ally toxic) transition metal catalysts needed. These metal contaminants have to be removed

    7

  • Reversible Addition-Fragmentation Chain Transfer Polymerization

    from the products, especially in industrial use and, due to their toxicity, limit the use of ATRP

    derived polymers. Recent developments led to improved methods to conduct ATRP poly-

    merizations. In the so called activator generated by electron transfer (AGET) ATRP process

    electron transfer is used instead of organic radicals to reduce the higher oxidation state tran-

    sition metal.38 This allows the use of ATRP catalyst systems in their more stable higher ox-

    idation states since these are reduced in situ before adding the initiator. In addition, the

    amount of transition metals could be reduced to ppm quantities using activators regene-

    rated by electron transfer (ARGET) ATRP.39 Reducing agents such as ascorbic acid or tin(II)

    2-ethylhexanoate can continuously regenerate CuI from CuII during the polymerization and,

    thus, ensure the catalytic amounts of transition metal needed.

    2.1.3 Reversible Addition-Fragmentation Chain Transfer

    Polymerization

    Reversible addition-fragmentation chain transfer (RAFT) polymerization was discovered at

    Australias CSIRO in the late 1990s.4042 In parallel, a similar polymerization technique,

    macromolecular design by interchange of xanthate (MADIX) was invented in France.14 Both

    RAFT and MADIX are based on an addition-fragmentation chain transfer mechanism.43,44

    While MADIX refers to polymerizations mediated by xanthates, RAFT is usually mediated by

    dithioesters or trithiocarbonates. Successful RAFT polymerization depends on the design of

    the RAFT agent. To date many different RAFT agents have been reported.40,41,4556 Both R-

    and Z-group allow for fine tuning of the performance of the RAFT agent. In principle the

    R-group should be a good leaving group and liberate a relatively stable free radical with the

    ability to initiate a polymerization. Therefore, the type and structure of the R-group can have

    tremendous impact on the polymerization kinetics and the overall control. In contrast, the

    Z-group is responsible for the ability of the C=S double bond to react with growing radicals

    and the mean lifetime of the resulting intermediate radical formed.

    The RAFT process is basically a free radical polymerization carried out in the presence

    of a particular chain transfer agent, the so-called RAFT agent. Accordingly, traditional

    methods to generate radicals from commercially available initiators such as 2,2-azobis(2-

    methylpropionitrile) (AIBN) or 4,4-azobis(4-cyanovaleric acid) (V-501) are used. The initia-

    tion step can also be carried out by photoinitiators or gamma irradiation.51,5759 Neglecting

    8

  • Reversible Addition-Fragmentation Chain Transfer Polymerization

    the solvent cage effect, the produced radical I can either add to a monomer to initiate a

    growing chain or add to the RAFT agent (Scheme 4). Due to the high transfer constants of

    the most RAFT agents it is unlikely that more than a few monomers add to a growing chain

    before the chain adds to a RAFT agent. Thus, via both pathways the intermediate radical is

    reversibly formed which then can either fragment to give the RAFT agent and the growing

    radical chain or cleave the R-group homolytically. The latter is the desired reaction pathway

    and requires the R-group to be a better leaving group than the oligomer or polymer chain

    and to be capable to initiate a polymerization. Hence, in the beginning of the polymeriza-

    tion, the so-called pre-equilibrium, all initial RAFT agents should be activated and converted

    into RAFT agents containing oligomeric or polymeric R-groups (macro-RAFT agents).

    Pre-equilibrium:

    Main equilibrium:

    Pi R

    PjPi

    Z

    SSR

    Z

    S SRPi

    Z

    S SPi+ +

    + +Z

    SSPj

    Z

    S SPjPi

    Z

    S SPi

    Scheme 4: Mechanism of the RAFT process.

    The number of polymer chains is controlled by the amount of RAFT agents and not by the

    number of initiator radicals produced. Once all RAFT agents are converted into macro-RAFT

    agents, the main equilibrium starts. Here the reversible addition-fragmentation reactions

    are dominant with a high amount of dormant chain ends compared to free radicals. In an

    ideal RAFT polymerization all polymers are initiated by the R-group of the RAFT agent. As a

    result the degree of polymerization is controlled by the ratio of monomer to RAFT agent and

    the molar mass at a certain conversion can be calculated according to:

    9

  • Reversible Addition-Fragmentation Chain Transfer Polymerization

    Mn,theor y =[Monomer ]0 MMonomer

    [C T A]0 +2f[I ]0 (1ekd t )+MC T A (2.1)

    where [Monomer]0 stands for the initial monomer concentration, MMonomer is the mole-

    cular weight of the monomer, is the conversion, [CTA]0 is the concentration of RAFT agent,

    f is the initiator efficiency, [I]0 is the initiator concentration, kd is the initiator decomposition

    rate constant, and MC T A is the molecular weight of the RAFT agent. If the ratio of [CTA]0/[I]0

    is high, initiation of polymer chains by initiator radicals can be neglected and equation 2.1

    simplifies to:

    Mn,theor y =[Monomer ]0 MMonomer

    [C T A]0+MC T A (2.2)

    The dithiobenzoate mediated RAFT polymerization involves two possible rate retarding

    effects, (i) an induction period during the initial stage with almost no polymerization activity,

    and (ii) rate retardation in the following phase with polymerization rates slower than in con-

    ventional radical polymerization systems.60 These effects depend on the concentration of

    RAFT agent. Rate retardation is only observed for polymerizations initiated by macro-RAFT

    agents. While the CSIRO team explained the rate retardation effect by slow fragmentation

    of the carbon-centered intermediate radical, Monteiro, Brouwer and Fukuda postulated a

    cross-termination of the intermediate radical with a growing radical chain.60

    In all theoretical models the retardation effect increases with increasing stability of the in-

    termediate radical. In dithiobenzoate mediated RAFT systems delocalization of the radical

    within the aromatic system is assumed to be responsible (Scheme 5).61

    Replacing the phenyl group by a benzyl moiety led to significantly lower rate retardation

    effects since the delocalization of the radical center is effectively suppressed.62 Further-

    more, para-substituted dithiobenzoates showed significantly lower rate retardation effects.63

    This observation indicates delocalized radicals and potential side reactions since the para-

    position is less prone to radical attacks whereas the stability of the intermediate radical

    should remain unchanged using the substituted RAFT agent. In order to investigate the

    mechanistic aspects of the RAFT process, radical storage experiments were applied to cumyl

    10

  • Reversible Addition-Fragmentation Chain Transfer Polymerization

    S SPjPi

    S SPjPi

    S SPjPi

    S SPjPi

    Scheme 5: Assumed resonance structures of the intermediate radical in dithiobenzoate mediated

    RAFT polymerizations.

    dithiobenzoate (CDB) mediated styrene and methylacrylate systems59,64 as well as to non-

    retarding RAFT agents.65 Systems containing dithiobenzoate RAFT agents are capable to

    store radicals for a certain time and, moreover, can induce a polymerization afterward with-

    out the need of initiators. However, such experiments cannot be used to identify the chemi-

    cal nature of these intermediates. In order to determine whether or not cross-termination

    occurs, electronspray ionization mass spectrometry (ESI-MS) was used. Until now, no star

    polymers could be detected by this technique.66,67 This lack of experimental evidence for

    cross-termination reactions is a reason for continued debates about the origin of rate retar-

    dation in the RAFT mechanism. 13C NMR measurements, in contrast, provided evidence

    about the existence of 3- and 4-arm star polymers when high initial concentrations of RAFT

    agents were used. This led to the assumption that the intermediate radicals undergo side

    reactions with growing radical chains.68

    On the theoretical level, the RAFT process can neither be fully explained using the cross-

    termination model nor the slow-fragmentation model. The cross-termination model is in

    good agreement with the experimentally observed kinetic aspects of the main equilibrium

    but predicts significant concentrations of termination products which could not yet be de-

    tected. On the other hand, the slow-fragmentation model is in agreement with the non-

    stationary polymerization rate in the pre-equilibrium and the experimentally observed ra-

    dical storage effects. However, the predicted intermediate radicals have not been observed

    in ESR spectroscopy and it is contradictory to the observations for the main equilibrium.

    A combined model which accounts for the differences in polymerization rate within the

    pre- and the main equilibrium and which also includes yet unknown mechanistic aspects

    of the RAFT process was developed by Buback who introduced the reversibility of the cross-

    termination reaction into the theoretical calculations.69 This led to good results and ex-

    plained the absence of star shaped polymers in the RAFT polymerizations. However, the

    11

  • Reversible Addition-Fragmentation Chain Transfer Polymerization

    reaction mechanism still needs to be verified and these side reactions do not explain that

    carbazoles show similar behavior to dithiobenzoates. Recently, Perrier et al. combined these

    two conflicting models assuming reversible termination of the intermediate radicals only

    for low molecular weight species but the slow fragmentation to be prominent during the

    main equilibrium.70 Termination of short intermediate radicals would explain the absence

    of 3-arm star polymers since the changes in molecular weight and hydrodynamic volume

    are negligible for low molecular weight species. Furthermore, matrix-assisted laser desorp-

    tion/ionization time-of-flight (MALDI-TOF) spectrometry experiments proved the existence

    of such short cross-termination products.71 The basis for this assumption is that cross-

    termination is diffusion controlled and, hence, small intermediates show much higher ter-

    mination rate coefficients. In a following publication the same group predicted these short

    intermediates to be maximum dimeric suggesting that cross-termination occurs only in the

    very early stages of the RAFT process.72

    In comparison to NMP and ATRP, RAFT polymerization is particularly robust and versatile

    and can be applied to many different monomer classes. However, the choice of RAFT agent,

    which commercial availability is limited, is crucial for successful polymer formation. More-

    over, the sulfur containing end groups often result in red to yellow-colored polymers which

    limits the application of RAFT in industrial processes without special treatment.

    2.2 Block Copolymer Self-Assembly

    Self-assembly of amphiphilic molecules into aggregates of different size and shape is

    mainly determined by intermolecular interactions such as van der Waals, hydrophobic-

    hydrophobic, hydrogen-bonding, or electrostatic interactions.73 Such systems are dynamic

    in nature and external changes (e. g., pH, temperature, concentration, etc.) can lead to

    changes of the formed aggregates. Both the thermodynamics of the self-organization pro-

    cess and intra-aggregate forces between molecules within the same aggregate determine the

    equilibrium structure formed. Considering the self-assembly of small amphiphilic lipids,

    two major forces govern the self-organization process, (i) hydrophobic attraction of the hy-

    drocarbon part and (ii) hydrophilic, ionic or electrostatic repulsion of the head groups. Thus,

    at a certain head group area (Figure 1) the energy of repulsive interactions reaches a mini-

    mum.

    12

  • Reversible Addition-Fragmentation Chain Transfer Polymerization

    a0

    v

    lc

    Figure 1: Schematic representation of an amphiphilic molecule with head group area a0, chain vol-

    ume v, and chain length lc .73

    The size of the formed aggregates is controlled by entropy which will favor the forma-

    tion of particles with the smallest aggregation number. While larger structures will be en-

    tropically unfavorable, smaller particles will suffer from increased repulsive interactions of

    the head groups and, hence, be energetically disfavored. For low molecular weight lipids the

    value of the packing parameter v/a0lc , where a0 is the optimal head group area, v is the hy-

    drocarbon volume, and lc is the critical chain length, determines the shape of the resulting

    aggregates (Figure 2).74

    Spherical micelles are formed only when the optimal head group area a0 is large in com-

    parison to the hydrocarbon volume v that the radius of the formed micelles does not exceed

    the critical chain length lc . Geometrical considerations for a spherical micelle with radius rm

    and aggregation number N give,

    N = 4r2m

    a0= 4r

    3m

    3v(2.3)

    which becomes

    r2m =3v

    a0(2.4)

    Hence, amphiphiles only assemble into spherical micelles if,

    v

    a0lc< 1

    3(2.5)

    With decreasing size of head group cylindrical micelles, bilayers, vesicles or inverted mi-

    celles are formed (Figure 2). Changes in temperature can cause changes in both a0 as well as

    lc depending on the amphiphile.

    13

  • Reversible Addition-Fragmentation Chain Transfer Polymerization

    Spherical Micelles

    Cylindrical Micelles

    Flexibel Bilayers or Vesicles

    Planar Bilayers

    InvertedMicelles

  • Reversible Addition-Fragmentation Chain Transfer Polymerization

    micelles, cylindrical micelles, and vesicles depending on the relative block length within AB

    diblock copolymers.

    100 nm

    100 nm

    100 nm

    Figure 3: Left: Schematically shown self-organization of diblock copolymers into spherical micelles,

    cylindrical micelles, and vesicles; Right: (A-C) cryoTEM images of aggregates formed by PB-b-PEO;

    (D-F), TEM images of aggregates formed by PS-b-PAA. A and D show vesicles; B and E show cylindrical

    micelles; C and E show spherical micelles.81

    Furthermore, two kinds of spherical micelles can be distinguished according to the rela-

    tive lengths of the blocks in an AB diblock copolymer, (i) star-like micelles with a small hy-

    drophobic core compared to the large corona, formed by the hydrophilic block, and (ii) crew-

    cut micelles with a large hydrophobic core and stretched short hydrophilic coronal chains

    (Figure 4). For simple amphiphilic block copolymers dissolution in a selective solvent for

    one block is the most straight forward method for the preparation of micelles. However, de-

    pending on the monomers used, non-equilibrium aggregates may be formed.82 In order to

    obtain micelles closer to a thermodynamic equilibrium, the block copolymers are often dis-

    solved first in a nonselective solvent followed by slow addition of a selective solvent for one

    block. The commercially available amphiphilic block copolymers from poly(ethylene oxide)

    and poly(propylene oxide) (PPO) are well-known to form micelles in aqueous solutions con-

    taining a PPO core surrounded by a dense PEO layer and dangling PEO chains forming the

    outer corona.83

    15

  • Reversible Addition-Fragmentation Chain Transfer Polymerization

    star-like crew-cut

    hydrophilic shell hydrophobic core

    Figure 4: Schematic representation of a star-like (left) and a crew-cut (right) micelle.

    In the case of diblock copolymers of polystyrene (PS) linked to PEO in aqueous solution,

    two populations with a hydrodynamic diameter Dh of 40 nm and 150 nm have been observed

    in TEM and light scattering studies.84 The authors concluded that the small aggregates are

    regular micelles while the large particles consist of loose clustered micelles. Further investi-

    gations using SANS, DLS and SLS proved the formation of anisotropic clusters of PS-b-PEO

    micelles in aqueous solutions of up to 10 wt%.8588 These clusters may be the result of merg-

    ing of initially formed micelles, as supported by controlled deaggregation of the clusters into

    micelles upon the addition of toluene or inorganic salts. The reason for cluster formation

    involves attractive interactions between the outer PEO chains such as hydrogen bonding or

    hydrophobic interactions.89

    The key problem of all block copolymer assemblies is how to control the structure and its

    dimensions by choosing appropriately the length of each block and, moreover, how to trig-

    ger structural changes such as transitions from spherical into cylindrical micelles by ex-

    ternal stimuli. Extensive theoretical and computational work indicates that especially the

    length of the hydrophilic corona forming block is crucial for the dimensions of the micelles

    formed.90,91 Besides controlling the dimensions of spherical micelles, it would be highly

    desirable to control the morphology of block copolymer aggregates with view on potential

    applications in nanotechnology. Pioneering work of Eisenberg et al. demonstrated that for

    PS-b-PAA block copolymers with a constant polystyrene block (DP = 200), a decreasing de-

    gree of polymerization (DP) of the poly(acrylic acid) (PAA) block from DP = 21 to DP = 4

    results in spherical, rod-like, vesicular, as well as crew-cut assemblies (Figure 5).5

    16

  • Reversible Addition-Fragmentation Chain Transfer Polymerization

    Figure 5: PS-b-PAA block copolymers forming a) spherical (PS200-b-PAA21), b) rod-like (PS200-b-

    PAA15), c) vesicular (PS200-b-PAA8), and d) crew-cut (PS200-b-PAA4) aggregates depending on the PAA

    content.5

    The morphology depends on three main factors, (i) the streching of the core forming

    blocks, (ii) the core-corona interfacial energy, and (iii) repulsion between the corona form-

    ing blocks. Changes in one of these three factors will result in thermodynamic instability and

    lead to rearrangement into thermodynamically more stable morphologies.

    Discher and Eisenberg proposed a general rule to predict the aggregate morphology

    from the block copolymer composition where fhydr ophi l i c is the relative content of the

    hydrophilic block:8

    Spherical micelles are formed when fhydr ophi l i c > 45%.

    Rod-like micelles are formed when fhydr ophi l i c < 50%.

    Vesicles are formed when fhydr ophi l i c ~ 35%.

    Inverted microstructures and crew-cut micelles are formed when fhydr ophi l i c < 25%.

    17

  • Reversible Addition-Fragmentation Chain Transfer Polymerization

    However, this rule has no universal validity. The chemical nature of the blocks used as

    well as the overall molecular weight of the block copolymers influence the aggregation in a

    way that is not yet fully understood.

    a) b)

    c)

    Figure 6: Cryo-TEM images of PEHA120-b-POEGA50-b-PFDA40 at 0.5 wt% in aqueous solutions a)

    dispersed at 25C; b) dispersed at 25and annealed for 21d at 78C; c) dispersed at 70C.92

    In addition, double hydrophobic ABC triblock copolymers containing two insoluble

    blocks assembled into core-shell-corona micelles with the two hydrophobic blocks form-

    ing the core and the shell and the hydrophilic block forming the solubilizing corona.9396

    For polymers with the hydrophilic block in the middle, the aggregation behavior can be as-

    sumed to be similar to ABA triblock copolymers as long the terminal blocks have a simi-

    lar degree of polymerization. In the case of double hydrophilic terpolymers, the insoluble

    block can be either in the middle or at one chain end. In the case of a terminal hydropho-

    bic block, core-shell-corona micelles are expected. In contrast, when the hydrophobic block

    is located in the middle of the block copolymer, non-centrosymmetric micelles can be ob-

    served due to the heterogenous corona.97,98 Recently, so-called multicompartment micelles,

    ABC triblock copolymers with a water-soluble shell and a segregated hydrophobic core have

    been published.99,100 The hydrophobic core was composed of hydro- and fluorocarbon

    containing blocks which showed segregation and the formation of two distinct domains.

    For instance, cryo-TEM measurements of poly(ethylhexyl acrylate)-b-poly(oligoethylene

    glycol monomethylether)-b-poly(tetrahydroperfluorodecyl acrylate) (PEHA120-b-POEGA50-

    b-PFDA40) showed the segregated dark fluorocarbon domains within the self-assembled

    18

  • Stimuli Responsive Block Copolymers

    micelles (Figure 6).92 Simultaneous, selective uptake of lipophilic and fluorophilic guest

    molecules into the corresponding hydrophobic polymer domains of multicompartment mi-

    celles was monitored by UV-vis101 and NMR spectroscopy.102,103 Besides spherical micelles

    and vesicular structures also helical104,105 or disk-like81 assemblies of block copolymers have

    been observed (Figure 7).

    Figure 7: A) Schematic illustration of self-assembled ABC block copolymers into disk-like micelles, B)

    TEM images displaying disk-like micelles of PAA-b-PMA-b-PS triblock copolymers (amine/acid ratio

    1/1) in a mixture of 40% water and 60% THF: (i) EDA as the counterion; (ii) EDDA as the counterion,

    C) (i) tilted TEM images of disk-like micelles (EDDA as the counterion, amine/acid ratio 0.3/1), (ii)

    CryoTEM images for the same sample where the disks are either parallel to the electron beam axis

    (arrows 1) or perpendicular to the electron beam axis (arrows 2). Scale bars = 200 nm.81

    2.2.1 Stimuli Responsive Block Copolymers

    In aqueous solution stimuli sensitive polymers are typically switched from a hydrophilic to a

    hydrophobic state or vice versa. Both physical (temperature, UV) and chemical (pH, redox)

    stimuli can be applied in order to change the hydrophilicity of polymers. In some cases poly-

    mers are even sensitive to more than one external stimuli. Poly(amine)s and poly(carboxylic

    acid)s undergo distinct changes in hydrophilicity upon protonation or deprotonation, re-

    spectively.106 Thermoresponsive behavior of polymers, in contrast, may be distinguished by

    19

  • Stimuli Responsive Block Copolymers

    two different types. Either polymers are soluble above a certain temperature, exhibiting an

    upper critical solution temperature (UCST), or they are soluble below a certain temperature,

    showing a lower critical solution temperature (LCST). In contrast to the UCST phenomenon

    which is caused by unfavorable enthalpy, the LCST is entropically driven.107 Rushbrooke

    suggested already in 1938 that intermolecular hydrogen bonding may cause the LCST phe-

    nomenon.108 In 1960 Freeman and Rowlinson reported that hydrocarbon polymers show a

    LCST in hydrocarbon solvents.109 This finding was validated by a universal theory of Flory

    on the LCST behavior of polymers in solution.110112 LCST behavior is widespread among

    non-ionic hydrophilic polymers,113116 and is attributed to the balance of polymer-solvent

    hydrogen bonding and polymer-polymer hydrophobic interactions.

    O

    O

    O

    O

    O

    O

    O O

    OO

    O

    OO

    O

    O

    O

    OO

    O

    O

    O

    O

    O

    O

    O

    OO

    O

    O

    O

    O

    O

    OO

    O

    H OH

    HOH

    HO H

    HO

    H

    HOH

    HO

    H

    HOH

    HO H

    H OH

    H OH

    HOH H

    OH

    HOH

    HOH

    O

    OO O

    O O

    O O

    OO O

    O

    O

    O

    O

    O

    O

    O

    O

    O O

    O OO

    OOO

    OO

    O

    O

    O

    O

    O

    O

    HO

    H

    H OH

    HO H

    HO H

    HO

    H

    HO

    H

    HO

    H

    HO

    H

    H OH

    H OH

    H OH

    H OH

    H OH

    H OH

    TLCST

    Figure 8: Phase transition of thermoresponsive polymers in water.

    Nevertheless, the exact position of such a phase transition can not be predicted by sim-

    ple hydrophilic-hydrophobic balance considerations.115,117,118 Moreover, the exact phase

    transition temperature of a polymer depends on additional factors such as molar mass, ar-

    chitecture, end groups, concentration or added salts.113,116,119,120 While the UCST increases,

    the LCST usually decreases with increasing molecular weight.121 In addition, an increase

    in meso diads decreases the LCST while an increase in racemo diads increases the LCST

    of poly(isoproylacrylamide) (PNIPAM).122,123 Isotactic PNIPAM was found to be more hy-

    drophobic than atactic one with phase transition temperatures of about 24C.124 Thus, tac-

    ticity of polymers seems to play an important role on their phase separation behavior. Also

    the heating rate may effect the measured phase transition temperature markedly.125 An in-

    crease of heating rate from 0.2 to 5 K/min can result in an increase in the apparent cloud

    point values of up to 10C due to insufficient heat transfer.121,126

    20

  • Stimuli Responsive Block Copolymers

    Figure 9: Thermodynamically stable random coil, crumpled coil, molten globule, and globule of

    homo-PNIPAM chains in aqueous solution during the thermally induced collapse.127

    The collapse of single PNIPAM chains in dilute solutions was intensively studied by

    several analytical techniques. Though thermosensitive homopolymers usually precipitate

    above their LCST, they show the formation of kinetically stable mesoglobules in highly dilute

    solutions ( 106 g mol1).128,129 Wu and coworkers studied the coil-to-globule and globule-

    to-coil transition using dynamic and static light scattering.128,130,131 After the coil-to-globule

    transition, which was observed by a decrease of Rg /Rh from 1.50 to 0.56, the average chain

    density decreased to 0.34 g/cm3 indicating that even the fully collapsed mesoglobules

    still contain about 66% water. The authors suggested, that between the coil and the glo-

    bule two other thermodynamically stable states exist, namely the crumpled coil and the

    molten globule127,130 (Figure 9). Winnik et al. used light scattering in combination with

    microcalorimetry for fluorescently labeled PNIPAM chains.129 Fluorescence spectroscopy

    indicated that PNIPAM mesoglobules undergo a gradual transition from fluidlike particles

    into hard spheres. Liu et al. reported on the observation of a two-stage transition of pyrene

    labeled PNIPAM chains by a combination of fluorescence and stopped-flow techniques. A

    first fast transition of randomly coiled PNIPAM into crumpled chains was followed by a slow

    collapse into compact globules.132 Also Aseyev and coworkers observed the formation of

    mesoglobules of PNIPAM, poly(N-vinyl caprolactam) (PVCL), and poly(vinyl methyl ether)

    (PVME) homopolymers.133 Although theoretical calculations suggest that homopolymers

    which form mesoglobular aggregates should assemble into cylindrical structures rather that

    21

  • Stimuli Responsive Block Copolymers

    spherical,134 no such experimental observation has been made so far. A combination of

    static and dynamic light scattering, instead, excluded cylindrical aggregates for PVCL, PNI-

    PAM, and PVME homopolymers but suggested spherical particles with relatively narrow size

    distributions.133

    The simplest examples of stimuli-responsive block copolymers contain one permanently hy-

    drophilic or hydrophobic block and one block which undergoes a phase transition from hy-

    drophilic to hydrophobic when external stimuli such as temperature, pH, ionic strength or

    UV-light are applied. Perrier et al. , for instance, synthesized diblock copolymers from PNI-

    PAM and poly(dimethylacrylamide) (PDMA) varying the length of the PDMA block.135 The

    authors stated that not only the relative length of the blocks determines whether micelles or

    larger aggregates are formed but also the absolute length of the hydrophilic block. Moreover,

    they demonstrated that the formed micelles are able to reversibly incorporate hydrophobic

    dye molecules, in this case the relatively large 2,6-diphenyl-4-(2,4,6-triphenyl-N-pyridino)

    phenolate also known as Reichardts dye (Figure 10). McCormick et al. synthesized block

    copolymers from PNIPAM and PDMA by aqueous RAFT polymerization at room tempera-

    ture and observed reversible micelle formation by passing the phase transition temperature

    of PNIPAM.136 An increase of the relative length of the PNIPAM block led to larger particles

    when the aqueous solutions were heated above 32-36C. Tenhu et al. studied diblock copoly-

    mers from PNIPAM with hydrophobic blocks of polystyrene or poly(tert-butylmethacrylate)

    obtained from RAFT polymerization.137 By variations of the length of the PNIPAM block sim-

    ilar observations as by Perrier et al. were made. Not only the relative length of the blocks is

    crucial for the formation of micellar aggregates. Longer PNIPAM chains led to larger aggre-

    gates which was interpreted in a way that the long PNIPAM chains destabilize the forma-

    tion of hydrophobic cores of polystyrene or poly(tert-butylmethacrylate). Interestingly, even

    at prolonged elevated temperatures of 50C for several days the formed particles remained

    stable and did not precipitate from solution. Marty, Destarac et al. synthesized PNIPAM-b-

    poly(butylacrylate) (PBA) diblock copolymers with varying length of the PBA blocks.138 With

    increasing PBA length the observed aggregates showed increasing hydrodynamic diameters.

    This was explained by an increasing aggregation number Nag g of the formed micelles. More-

    over, the PBA block led to a decrease of the phase transition temperature of the PNIPAM

    block by about 6C. The cloud point was decreased even further when the length of the PBA

    block increased. Complementary results were obtained by Armes et al. for AB block copoly-

    22

  • Stimuli Responsive Block Copolymers

    mers from methyl vinyl ether (MVE) and methyl triethylene glycol vinyl ether (MTEGVE).139

    The cloud points of the block copolymers increased from 18C to 84C with increasing length

    of the more hydrophilic polyMTEGVE block.

    Figure 10: Bathochromic and hypsochromic shift of Reichardts betaine dye in an aqueous solution of

    PDMA58-b-PNIPAM61 at a concentration of 1 w% upon heating and cooling between r.t. and 55C.135

    Especially block copolymers from PNIPAM and PEO found much attention through-

    out the recent years, because of the lower critical solution temperature of PNIPAM (32C)

    close to human body temperature (37C) and the biocompatibility of PEO.113,140143 Such

    block copolymers are expected to be interesting candidates for drug delivery and biomed-

    ical applications.115,140,144149 For instance, Feijen et al. synthesized PEO-b-PNIPAM block

    copolymers which self-assembled into spherical micelles in aqueous solution once the tem-

    perature was increased above the cloud point of the PNIPAM block at about 31C.141 Hennink

    et al. studied the self-aggregation of PEO-b-PNIPAM block copolymers in dilute aqueous so-

    lution, too.150 With increasing PNIPAM chain length the particle size was found to decrease.

    These findings were interpreted in terms of increased dehydration of the thermosensitive

    block which results in more densely packed hydrophobic cores. These observations are in

    direct contrast to the results obtained by Marty and Desterac,138 who found larger particles

    with increasing length of the hydrophobic block (vide supra). Moreover, the heating rate ap-

    peared to influence the size of the formed aggregates with a fast heating protocol resulting

    in smaller particles than observed for slow heating rates. In addition, the particles formed

    by fast heating had lower polydispersity indices than the ones from slow heating. Based on

    these finding the authors tested a heat shock protocol. Relatively small amounts of polymer

    solution below the cloud point temperature were added to water at 40C, thus well above

    23

  • Stimuli Responsive Block Copolymers

    the LCST of PNIPAM. Such rapid heating led to even smaller particles with a Dh of 50 nm

    and very low polydispersities around 0.04. Pispas et al. observed the same trend for dif-

    ferent heating protocols of PEO-b-PNIPAM diblock copolymers.125 Moreover, Shi et al. ob-

    served a concentration dependence of micellar size and size distribution of PEO-b-PNIPAM

    diblock copolymers. Surprisingly, higher concentrations led to smaller and more narrowly

    distributed aggregates than smaller concentrations at which loose micellar assemblies or

    even clusters appeared.142

    More dynamic aggregation behavior was observed for double responsive diblock copoly-

    mers. By the use of orthogonal stimuli or by the combination of one LCST and one

    UCST block, so called schizophrenic block copolymers can be obtained in which micelles

    and reverse micelles can be selectively formed according to the external stimulus applied.

    For instance, double thermoresponsive block copolymers from PNIPAM and poly(3-[N-(3-

    methacrylamidopropyl)-N,N-dimethyl]ammoniopropane sulfonate) (PSPP) having a LCST

    and a UCST block, respectively, have been synthesized.151 At low temperatures PSPP-core

    PNIPAM-shell micelles are obtained, at intermediate temperatures the block copolymers

    are molecularly dissolved whereas at high temperatures reverse micelles with a PNIPAM

    core and a PSPP shell are formed. Similar systems have also been reported by Maeda and

    Armes.152,153 In addition, schizophrenic diblock copolymers sensitive to changes in pH val-

    ues have been reported. A zwitterionic block copolymer of poly(4-vinylbenzoic acid) (PVBA)

    and poly(2-N-(morpholino)ethyl methacrylate) (PMEMA) showed PVBA-core micelles be-

    low pH 6. Increasing the pH above 6 led to dissolved polymers which aggregated into

    PMEMA-core micelles at elevated temperature or in the presence of Na2SO4.154 In addition,

    even triple responsive diblock copolymers exhibiting a third redox155 or sugar response156,157

    were published recently but especially the redox switch is often not reversible. Such

    schizophrenic block copolymers are especially interesting due to their ability to aggregate

    into different structures under changing conditions. Double responsive diblock copolymers

    with a thermo- and a pH-responsive block from PNIPAM and poly(N,N-diethylaminoethyl

    methacrylate) (PDEAEMA), respectively, could even be switched from micellar into vesicular

    aggregates in response to changes of the external conditions158 (Figure 11).

    24

  • Stimuli Responsive Block Copolymers

    a)

    b)

    pH

    pH

    Temp

    Temp

    Micelles inverse Micelles

    Micelles Vesicles

    Figure 11: Schizophrenic aggregation behavior of (a) PDEAEMA98-b-PNIPAM209 and (b)

    PDEAEMA98-b-PNIPAM392. Adapted from ref.158

    Beside diblock copolymers, multi responsive terpolymers with ABA or BAB sequence,

    containing a responsive block and a permanently hydrophilic B block, are well-known for

    their rich aggregation behavior. While BAB triblock copolymers often form spherical mi-

    celles with a core forming A block and corona forming B blocks, the situation becomes more

    complex for ABA systems. Theoretical discussion of the formation of flower-like micelles

    in which the dissolved middle block has to loop in order to incorporate both hydropho-

    bic A blocks into the core were done by Tirrell and coworkers.159 Calculations suggested

    that only polymers with very short hydrophobic A blocks will form flower-like micelles while

    for longer A blocks mixed star-/flower-micelles, with partially free dangling chain ends, are

    more probable. The assumption of flower-micelles formed by ABA triblocks was experimen-

    tally supported by the findings that the investigated aggregates had dimensions of those

    formed by diblock copolymers of half the molecular weight.159,160 Moreover, when the re-

    lative block length of the middle block was decreased below the size of the hydrophobic A

    blocks, no aggregation was observed. Likewise to their diblock counterparts, changes in size

    and morphology of triblock copolymer aggregates can be induced by changes of the solvent

    qualities or upon introducing stimuli-sensitive blocks. Wang et al. synthesized poly(stearyl

    methacrylate)-b-PNIPAM-b-poly(stearyl methacrylate) (PSMA-b-PNIPAM-b-PSMA) triblock

    copolymers which showed morphological transitions from vesicular into micellar structures

    with increasing water content in THF/water mixtures.161

    25

  • Stimuli Responsive Block Copolymers

    ABC triblocks with stimuli-sensitive blocks have been described by Eisenberg and cowork-

    ers who reported on the self-assembly of PAA26-b-PS890-b-P4VP40 as function of pH in

    DMF/THF/H2O mixtures.162 At low pH, vesicles with the P4VP blocks forming the outer shell

    and PAA blocks forming the inner shell were formed. At intermediate pH values, spherical

    and ellipsoidal aggregates were found while at high pH again vesicles, this time with outer

    PAA and inner P4VP block, were observed. Thus, an interconversion of vesicles with either

    outer PAA or P4VP blocks was possible by changing the pH of the solution (Figure 12).

    Figure 12: Schematic representation of the pH induced inversion of vesicles formed from PAA26-b-

    PS890-b-P4VP40 triblock copolymers.162

    P2VP58-b-PAA924-b-PBMA48 terpolymers also showed morphological changes at differ-

    ent pH values and temperatures in aqueous solutions.163 At pH 1, changes in tempera-

    ture led to either mesoglobules (T < 20C) or swollen micelles (T > 20C). An increase in

    pH from 8 to 11 and then to 12 led to spheres, toroidal nanostructures and finally to mi-

    crogels, respectively. Another example for double responsive ABC triblock coplymers is

    the aggregation of PEO-b-P4VP-b-PNIPAM in aqueous solution.164 At 25C a unimer-to-

    micelle transition occured when the pH was increased from 2 to 6.5. At elevated tem-

    peratures, above the cloud point of the PNIPAM block, micellar clusters could be con-

    verted into core-shell-corona micelles upon increasing the pH. Interestingly, even mor-

    phological changes from spherical into worm-like micelles were reported for double hy-

    drophilic ABC triblock copolymers from PEO114-b-PBA250-b-PDEAM135 (poly(diethyl acryl-

    amide) (PDEAM)), upon changes of the solution temperature165 (Figure 13). In addition,

    schizophrenic aggregation behavior of ABC triblock copolymers could be obtained from

    poly(2-(diethylamino)ethyl methacrylate)-b-poly(2-(dimethylamino)ethyl methacrylate)-b-

    poly(2-(N-morpholino)ethyl methacrylate) (PDEA-b-PDMA-b-PMEMA) by changes in pH

    and addition of Na2SO4.166

    26

  • Stimuli Responsive Block Copolymers

    Figure 13: Worm-like and spherical micelles formed from double hydrophilic ABC triblock copoly-

    mers depending on the solution temperature.165

    At pH 7.6 PDEA-core, PDMA-shell, PMEMA-corona micelles were formed which inversed

    to PMEMA-core, PDMA-shell, PDEA-corona micelles when Na2SO4 was added. Finally, ABC

    terpolymers can also form helical105 or hamburger-like167 superstructures under proper sol-

    vent conditions.

    Consequently, triple responsive ABC triblock copolymers are assumed to provide even more

    opportunities to control the self-aggregation and, moreover, trigger morphological changes

    stepwise. However, until now only very few examples of triple responsive terpolymers have

    been reported, all of which comprising three temperature responsive blocks. Aoshima

    et al. reported on the synthesis and aggregation behavior of triple thermoresponsive block

    copolymers from 2-ethoxyethyl vinyl ether (cloud point 20C), 2-methoxyethyl vinyl ether

    (cloud point 41C) and 2-(2-ethoxy)ethoxyethyl vinyl ether (cloud point 64C) obtained

    from sequential living cationic polymerization.168 Such ABC polymers showed a multi-

    step aggregation process from unimers to micelles followed by physical gelation and finally

    precipitation upon increasing the temperature. More recently, Zhu et al. reported on the

    multi-step phase transitions of triple thermoresponsive terpolymers in dilute aqueous so-

    lutions and studied their self-assembly by a combination of dynamic and static light scat-

    tering, NMR spectroscopy as well as ultrasensitive differential scanning calorimetry.169,170

    ABC triblock copolymers with thermoresponsive blocks from poly(N-propylacrylamide)

    (PNPAM, cloud point 22C), poly(N-isopropylacrylamide) (cloud point 32C) and poly(N,N-

    27

  • Stimuli Responsive Block Copolymers

    ethylmethylacrylamide) (PNEMAM, cloud point 56C) showed a multi-step self-organization

    from unimers into micelles or micellar clusters depending on the chain length of the

    poly(N,N-ethylmethylacrylamide) block, i. e., only a long poly(N,N-ethylmethylacrylamide)

    block was able to stabilize micellar aggregates. Whereas the influence of the chain length

    of the terminal high-LCST block on the aggregation process has been studied, the effect

    of changing the block sequence has not been explained. However, although the triblock

    copolymers PNPAM124-b-PNIPAM60-b-PNEMAM44 showed three changes in transmission

    intensity with increasing temperature, dynamic light scattering displayed only one thermal

    transition for homo-, di-, and triblock copolymers (Figure 14).

    Figure 14: Turbidity (left) and dynamic light scattering (right) measurements of PNPAM124-b-

    PNIPAM60-b-PNEMAM44 .169

    After passing the cloud point of the PNPAM block at about 24C, aggregates with a Dh

    of ~ 150 nm were observed for PNPAM124-b-PNIPAM60-b-PNEMAM44. Further heating led

    to a gradual decrease of Dh without any indication of a second pronounced change. Varia-

    tions in the relative block length, PNPAM124-b-PNIPAM80-b-PNEMAM44 and PNPAM124-b-

    PNIPAM80-b-PNEMAM160, then led to better results and changes of the hydrodynamic dia-

    meter with temperature became visible upon heating their aqueous solutions from 15-70C

    (Figure 15).170

    28

  • Stimuli Responsive Block Copolymers

    Figure 15: Hydrodynamic diameter (top) and apparent molar mass (bottom) of PNPAM124-b-

    PNIPAM80-b-PNEMAM44 () and PNPAM124-b-PNIPAM80-b-PNEMAM160 () as function of tempe-rature.170

    While the dynamic light scattering curves show more than a single transition during the

    heating process the interpretation and proposed aggregation pathway (Figure 16) is ques-

    tionable at least for the polymer with the shorter PNEMAM block . The smallest observed

    aggregates have a Dh of about 160 nm while its contour length is 62 nm. Thus, even in the

    fully extended state no single micellar aggregates with a hydrodynamic diameter of mini-

    mum 160 nm can be formed.

    29

  • Stimuli Responsive Block Copolymers

    Figure 16: Proposed aggregation pathway of PNPAM124-b-PNIPAM80-b-PNEMAM44.169

    2.3 Alternating Copolymerization

    Nylon 6,6 is maybe one of the most famous alternating copolymers due to the wide range

    of commercial applications. Nylon 6,6 can be obtained by polycondensation of the cor-

    responding diamines and dicarboxylic acids. In principle, the polycondensation or step-

    growth polymerization of A-A and B-B monomers will always result in alternating copoly-

    mers since each monomer cannot react with itself. Thus, crosspropagation is the only pos-

    sible polymerization pathway. In chain growth systems such as a common radical copoly-

    merization of two monomers A and B, however, four different chain propagation reactions

    can take place, namely A as well as B can add either to an active chain of A or B. These four

    propagation pathways are governed by the propagation rate constants kab2 , kaa2 , k

    bb2 and k

    ba2 .

    The resulting composition of the formed copolymers is then determined by the ratios of the

    corresponding propagation rate constants

    = kab2

    kaa2; = k

    bb2

    kba2(2.6)

    Accordingly, if a copolymerization of A and B is started in a molar ratio of B/A, the molar

    ratio within the resulting polymer b/a can be calculated according to:171

    30

  • Alternating Copolymers of Maleic Anhydride and Styrene

    b

    a= B

    A B + A

    B + A (2.7)

    Alternating copolymers are usually only obtained by radical polymerization, if the

    monomer pair is forming charge-transfer complexes (CTC) which then homopolymerize, or

    if the crosspropagation rate constants are much higher than the homopolymerization rates.

    N-Vinylcarbazole, for instance, was reported to form 1:1 alternating copolymers with fuma-

    ronitrile and diethyl fumarate.172,173 Also alternating copolymers of styrene and acrylonitrile

    were synthesized by free radical polymerization in the presence of zinc chloride.174 In the

    special case of styrene (A) and maleic anhydride (B), the latter does not undergo homopoly-

    merization and, in addition, the propagation rate for styrene terminated polymers is much

    higher toward maleic anhydride than styrene monomers. As a result, kbb2 becomes zero, thus,

    also becomes zero and equation 2.7 simplifies to:

    b

    a= 1+ 1

    A

    B(2.8)

    2.3.1 Alternating Copolymers of Maleic Anhydride and

    Styrene

    Since the 1940s the copolymerization of styrene and maleic anhydride (MAn) is known

    to proceed in an alternating fashion.171,175 However, there is still disagreement about the

    origin of the alternating nature of this monomer pair. Mainly two, incompatible, expla-

    nations are discussed in the literature.176 On the one hand the alternation is claimed to

    arise from charge-transfer-complexes which show higher reactivity against the growing ra-

    dical than uncomplexed monomers.177,178 On the other hand the alternating copolymer-

    ization of styrene and maleic anhydride is discussed in terms of different reactivities of the

    monomers resulting in marked preferences for crosspropagation reactions.179 Sometimes

    it is argued that this general tendency is enhanced by donor-acceptor interactions between

    the growing radical chain ends and adjacent monomers. In the case of maleic anhydride

    and N-substituted maleimides with styrene, the very low tendency of maleic anhydride and

    31

  • Alternating Copolymers of Maleic Anhydride and Styrene

    N-substituted maleimides toward homopolymerization and the extremely favored crosspro-

    pagation with styrene can be accounted for the alternating nature of the polymerization pro-

    cess.180,181

    However, the former is still an issue of discussion since UV and NMR measurements proved

    the existence of styrene-maleic anhydride donor-acceptor complexes under usual polymer-

    ization conditions.177 In solvents such as methyl ethyl ketone or DMSO competitive com-

    plexation of the monomers by the solvent molecules takes place. However, the alternat-

    ing feature of the copolymerization alone cannot prove the involvement of CTCs. For in-

    stance, the relative reactivity of styrene and MAn with donor-type radicals, e. g., cyclohexyl

    radicals, was tested and showed that MAn was 850 times more reactive than styrene.182

    Likewise, acceptor type radicals, benzoyloxy radicals, reacted 50 times faster with styrene

    than with MAn.183,184 Thus, the alternation can also arise from the different reactivities

    of the monomers favoring crosspropagation. Moreover, the question has to be addressed

    whether the CTCs have a higher reactivity toward a growing polymer chain than uncom-

    plexed monomers since this cannot be assumed a priori. In fact, theoretical calculations are

    difficult for this problem since the structure of the formed styrene-MAn complex can only

    be assumed from 1H NMR data but has large impact on the potential reactivity. Theoret-

    ical calculations, however, have shown that CTCs may indeed participate in the polymer-

    ization but only one reaction proceeds faster for the complex, namely the addition to an

    active styrene end. Electron spin resonance (ESR) spectroscopy as well as trapping of ac-

    tive radical chain ends was suggested to give further information about the polymerization

    mechanism. Indeed, by trapping styrene-MAn mixtures with 2-methyl-2-nitrosopropane at

    50C only styrene terminated molecules could be found in agreement with a polymerization

    of CTCs which add with their MAn side to the active styrene end.185 However, the trapping

    agent used is not very efficient in trapping maleic anhydride radicals and, therefore, the ob-

    servation made is not a proof for the proposed mechanism.176

    The latter, in contrast, is accounted for by the penultimate unit model (PUM).186 Here not

    only the terminal unit determines the crosspropagation rate constant but also the penul-

    timate unit (ki i i = k j i i ). Especially the decreasing content of alternation with increasingtemperature was often used to support the complex participation model (CPM). However,

    copolymerizations in the absence of CTCs are known to proceed more randomly at higher

    temperatures and the temperature dependence of copolymer composition vs. monomer

    32

  • Copolymerization of N-substituted Maleimides with Styrene

    feed can be described by both models (CPM and PUM). Thus, in order to rule out one of

    these two models other experimental data are required. Determinations of average propaga-

    tion rate coefficients as function of comonomer feed compositions using pulsed laser poly-

    merization were used by Klumperman et al. for the MAn-styrene system.187 The observed

    increase of the rate coefficients with increasing MAn fraction could be explained using the

    PUM but not by the CPM. Reductive demercuration was used to generate benzyl radicals in

    N-phenylmaleimide (NPhM) styrene mixtures in order to investigate the concerted addition

    of a styrene terminated polymer to a CTC of a N-substituted maleimide and styrene.188 The

    results showed almost exclusively single addition of styrene ruling out a concerted addition

    of a 1:1 maleimide-styrene complex. Further indication for the absence of CTC in RAFT-

    mediated styrene-MAn copolymerizations were reported using in situ 1H NMR spectroscopy

    during the initiation period.189 The results showed that, depending on the structure of the

    RAFT agent used, selective addition to a single monomer unit of either MAn or styrene takes

    place. Within the early stages of the copolymerization only one monomer unit adds to the

    growing polymer chain before termination takes place. In the case of complex participation,

    however, the styrene-MAn couple should be added to the chain end.

    2.3.2 Copolymerization of N-substituted Maleimides with

    Styrene

    Compared to maleic anhydride, the copolymerization of N-substituted maleimides with

    styrene is not strongly alternating190192 since the styrene content can be varied depending

    on the monomer feed.176 This aspect is also important for the discussion of CTCs involved in

    the polymerization mechanism since styrene rich copolymers can arise only from incorpora-

    tion of non-complexed monomers. Moreover, in contrast to MAn, N-substituted maleimides

    can be homopolymerized by anionic as well as radical polymerization.193196

    Many different N-substituted maleimides have been used as comonomers in order to in-

    crease the solubility and thermal stability,196,197 or functionalize polymer chains.180,181 Re-

    cent examples by Lutz et al. showed that the monomer sequence can be conveniently con-

    trolled even in radical polymerization when the kinetics are well known.180,181,198200 The

    sequential addition of various different maleimides to a growing styrene chain led to prepro-

    grammed distributions of functional side chain groups and up to four different maleimides

    could be incorporated along one polystyrene chain (Figure 17).

    33

  • One-Step Synthesis of Diblock Copolymers

    This concept works because the maleimides are incorporated immediately after addition

    and result in very narrow domains of alternating copolymer parts.

    N OO

    R1

    N OO

    R2

    N OO

    R3

    N OO

    R4

    t0 t1 t2 t3

    convS ~ 0.25 convS ~ 0.50 convS ~ 0.75

    Living chain-growth

    Controlled monomer addition

    Figure 17: General concept for the sequential addition of various functional maleimides to the poly-

    merization of styrene by ATRP according to Lutz et al. 201

    2.3.3 One-Step Synthesis of Diblock Copolymers

    In comparison to commonly applied sequential synthesis of block copolymers by con-

    trolled radical polymerization techniques, the one-step formation of block copolymers

    would be a significant advantage. The aforementioned alternating copolymerization of

    styrene and MAn or N-substituted maleimides is a promising candidate within this con-

    text. Recently a few reports were published on the one-step formation of AB diblock

    copolymers. The NMP polymerization of a 9:1 mixture of styrene and MAn as reported

    by Hawker et al. produced a poly(styrene-co-MAn)-b-polystyrene (P(S-co-MAn)-b-PS) block

    copolymer.202 During the initial phase of the polymerization the much faster crosspro-

    pagation takes place until MAn is consumed, followed by the homopolymerization of

    styrene. However, NMP did not result in an alternating block but in a copolymer with

    a 2:1 content styrene:MAn. By the use of RAFT, Li et al. obtained P(S-alt-MAn)-b-PS in

    a one-step synthesis.203 Hydrolysis of the MAn units gave an amphiphilic block copoly-